Skip to main content

Full text of "Entanglement engineering of one-photon wavepackets using a single-atom source"

See other formats


Entanglement engineering of one-photon wavepackets using a single-atom source 



00 
On 
On 

X> ■ 
in : 

(N . 
(N ' 

> : 
o . 
\o ■ 
o ■ 
<n : 
o . 

00 ■ 
On 1 

^— > ■ 

G ■ 

a ; 

3 : 



K.M. GheriW, C. Saavedra^ 1 * 2 ), P. ToiW 1 ), J. I. ChW 1 ' 3 ), and P. Zollert 1 ' 3 ) 
(1) Institut fur Theoretische Physik, Universitdt Innsbruck, Technikerstrasse 25/2, A-6020 Innsbruck, Austria 
(2) Departamento de Fisica, Universidad de Conception, Casilla 4009, Conception, Chile 
(3) ITP, University of California, Santa Barbara, CA 93106-4030 
(February 1, 2008) 

We propose a cavity-QED scheme for the controlled generation of sequences of entangled single- 
photon wavepackets. A photon is created inside a cavity via an active medium, such as an atom, and 
decays into the continuum of radiation modes outside the cavity. Subsequent wavepackets generated 
in this way behave as independent logical qubits. This and the possibility of producing maximally 
entangled multi-qubit states suggest many applications in quantum communication. 



Pacs number(s): 03.67.Hk, 03.67.-a 

Sources offering a great variety of entangled states are 
required for the implementation of many quantum com- 
munication and computation protocols jl],^|. With quan- 
tum communication || in mind the choice of photons as 
qubits is especially appropriate, since they can be eas- 
ily transfered over long distances. The standard source 
presently used in the lab is parametric downconvcrsion 
in a crystal |^]||. It is a reliable source of entangled 
twin-photons but the process is random and largely un- 
tailorable. Moreover, in practice its capability of gener- 
ating entanglement is limited to states comprising only 
two photons. In this Letter we propose a scheme for the 
controlled generation of many entangled photonic qubits. 
Our source of entanglement produces a train of single- 
photon wavepackets which are well resolved in time. This 
permits us to regard them as individual qubits. In its 
most simple implementation the setup consists of a sin- 
gle multilevel atom inside an optical resonator |^^] . The 
individual wavepackets are generated by applying an ex- 
ternal laser pulse to the atom prepared in a superposition 
state of its internal states. The coupling of the atom to 
the resonator allows the transfer of a single photon to 
the resonator and therefrom via cavity decay to the con- 
tinuum of radiation modes outside the resonator (possi- 
bly coupled to an optical fiber). An encoding of quan- 
tum information in the one-photon wave-packets could 
either take place by identifying two orthogonal polariza- 
tion states of the single photon with logical "0" and "1" , 
or by regarding the absence of a photon as logical "0" 
while its presence would correspond to logical "1". 

Our scheme offers a twofold advantage over already 
existing sources of entangled single-photon wavepackets 
such as down-conversion. It provides excellent control 
over the instances in time when a qubit is created as well 
as over the spectral composition of the wavepacket. The 
qubits may thus be generated with a well defined tact 
frequency and pulse shape. Moreover, repeated coherent 
recycling of the state of the atom after the generation of 
a photon wave packet gives rise to higher order entangle- 
ment between subsequent photons H . In this regard our 
scheme generalizes and extends recent work on sources of 
single photon wavepackets, commonly referred to as pho- 



ton guns (9) or turnstile devices |l(J by allowing the gen- 
eration of entangled multiphoton states. In particular, 
states such as the the three-particle GHZ state, and more 
generally n-qubit maximally entangled states (MES) can 
be generated. Since the individual wavepackets do not 
overlap in time, each can be sent to a different receiver 
node using simple classical gating operations. Such high- 
order entangled qubit states have immediate application 
in quantum cryptography |TT| and teleportation ]l2[ ], as 
well as in tests of non-locality and multiparticle inter- 
ference [ [[3) . Because entangled states of more than two 
qubits can be generated in a straightforward manner, our 
scheme is especially useful for quantum communication 
between many parties Jwj. 

Whilst the theory underlying our proposal can be for- 
mulated in a model-independent fashion it is more in- 
structive to illustrate the basic ideas using a specific 
model: we consider a single atom or ion trapped inside 
a cavity fflfy. For the atom we assume a double three- 
level A structure in the large detuning limit as depicted in 
Fig. 0. The levels \i a ) ( a = 0,1) are coupled to the upper 
levels \r a ) via classical fields Q, a {t)e~ l ^ at+ ^ a ^\ where 
uj a are the field center frequencies and the subscript refers 
to the two polarization states. The external control pa- 
rameters are the real amplitudes fl a (t) and the phases 
4> a {t). The levels |/ Q ) are coupled to the upper levels 
by the cavity modes a a (common frequency lu c but or- 
thogonal polarization), with coupling constants g a . The 
large detuning (S) assumption allows us to adiabatically 
eliminate the upper atomic levels. We are left with two 
two-level systems describable by generalized spin oper- 
ators Oi a j a — \i a ){ja\- The center frequencies of the 
external laser pulses fulfil the Raman resonance condi- 
tion. Note that any offsets can still be accommodated 
in the phases 4> a (t). The field outside the resonator is 
described by a continuum of harmonic oscillators with 
creation and annihilation operators b^ a (uj), b a (u>), respec- 
tively. The reservoir modes satisfy standard bosonic com- 
mutation relations: [b a (u>),bp(v)] — 5 a p8(u) — v). The 
coupling of the cavity and reservoir modes is assumed to 
be fl at aro und the cavity resonance frequency and equal 
to ©■ 



1 



Jr.) j£i»> 
~%) *® lie) |S) 




FIG. 1. A single atom with six internal states interacts 
with two cavity modes of orthogonal polarization ao, ai. In a 
Raman process (step 1) an initial superposition state of lev- 
els |io) and is transformed into an entangled cavity-atom 
state. Due to cavity leakage the photon will leave the cavity 
and produce a photon wave packet in the continuum modes 
outside the resonator. In step 2 the atom is recycled back to 
\io) and Between two photon generations levels |io) and 
can be coupled (step 3) to tailor the outgoing state. 

Hence the total system consists of three building 
blocks: the continuum outside the resonator, the cavity 
modes and the internal degrees of freedom of the atom 
inside the resonator. We switch to an interaction picture 
with respect to the free dynamics of the compound sys- 
tem. This eliminates the fast optical timescales from the 
dynamics and leaves us with a simpler Hamiltonian: 



H(t) = £ (i^ J duo (aabiiCuy^ - H. c.)) + V(t), 
V(t) =^fe(i)a iaiQ + \g a \ 2 aia a a fafa 



Pa(t) (c 



We have introduced the following abbreviations Q, a (t) = 
Q a (t)/2-y/8, g a = g a /VS, r a (t) = g a Cl a (t) and Q = 
cu + lu c . The time and intensity dependent terms in V(t) 
correspond to ac-Stark shifts arising from the adiabatic 
elimination of the upper atomic levels. 

We assume an atom initially prepared in a superposi- 
tion state \ip(0)) a = co|irj) + c iKi) which we wish to map 
onto a superposition of continuum (reservoir) excitations. 
The cavity and the reservoir modes are in their vacuum 
states. Since the dynamics contain no polarization mix- 
ing terms we may independently consider those degrees 
of freedom corresponding to a single index a. We may 
thus work with a smaller system and intermittently drop 
the index a. The final state of the total system can be ob- 
tained using the superposition principle and issuing each 
partial solution with the appropriate probability ampli- 
tude c a . As a starting point we will discuss briefly the 
generation of single photon wavepackets entangled with 
the atom. Since there can be at most a single excitation 
transfered to the continuum we are led to the following 
ansatz for the state of the total system: 



\^{t)) = |^(t))ac|0) r + / dw|¥»o(t))ac|W- 



(1) 



Here |ys(i))ac and \(p(t)) ac denote atom-cavity states 
with and without the transfer of a photon having taken 



place into the reservoir mode with frequency Co, respec- 
tively. Note that |y>(t)) ac describes the atom-cavity state 
before the photon has been lost to the reservoir: 

|^W>ac = a(t) e - ie W| l )|0) c + C / We-^l 2t |/)|l) c , 

where 9(t) = J* dt'fl 2 (t'). Applying a to this state 
projects the coupled atom-cavity system into the state 
|/)|0) c which is not coupled by V. We thus find: 



|<Aa(*))ac = 



dt 



a\y(t')) s 



(2) 



If we insert this expression into the equation for | (/?(£)) ac 
and perform the Markov approximation we arrive at a 
simple closed equation: 



-(K c a^a + iV(t))\ip(t)) a 



(3) 



This now permits us to specify the sought evolution equa- 
tions for the coefficients in the ansatz for |y>(t)) ac : 

Ci(t) = -r(t)Cf(t)exp(iO e (t)), 

C f (t)= f dt'r(t')a{t')eM-(K c (t-t / )+ i6 c {t'))), 
Jo 

where 9 c (t) = 9(t) + <j)(t) — \g\ 2 t. In the limit of an over- 
damped cavity the integral will get a non-zero contribu- 
tion only from those times t' which are close to t on the 
scale of the cavity lifetime k" 1 . To good approximation 
it thus holds that 



where fi(t) = r 2 (t)/n c , and jii(O) = 0. The actual ob- 
ject of interest is the state of the continuum of radiation 
modes outside the cavity. We thus insert the above result 
for Cf into Eq. (0) and find: 



IM*)) 



(4) 



G(u,t)a fi \i)\Q) t 



Eq. (|J) indicates a direct mapping of the initial atomic 
state to the final one accompanied by the creation of a 
wavepacket with spectral envelope G(u>,t), cf. step 1 in 
Fig. 0. To make the scheme practical two constraints 
have to be kept in mind. First of all we would like to im- 
plement an efficient transfer of the photon to the contin- 
uum. Secondly, we are only interested in pulse sequences 
that terminate after a finite time T> k^ 1 . This would 
be warranted if at the time t — T the atom is with near 
certainty in its internal state |/), i.e., iff /i(T) 3> 1. Re- 
calling the definition of n(i) this sets for any given pulse 
duration T a lower bound for the minimum size of the 
pulse area of the applied laser field. Under this assump- 
tion the total system state for times t > T is given by: 



2 



\m) = f^ Cct i?t(o,r)|/ Q )j|o>.|o> Co |o) Cl , (5) 

where B^ (tj , T) is the creation operator of a one-photon 
wavepacket with logical or polarization state a within the 
time window from tj to tj + T: 

Bl{tj,T) = J dve^G^TpjQ). (6) 

Note that the spectral envelope now carries a subscript as 
the system parameters need not be the same for each of 
the effective two-level systems we use to implement the 
mapping. In brief, we have shown how one can trans- 
form an initial atomic superposition state into an entan- 
gled atom-continuum state. Had we used only a single 
A system we would have recovered the photon gun || , a 
tailorable simple single-photon source. The novel aspect 
here is the residual entanglement between the internal 
atomic state and the polarization state of the outgoing 
wavepacket. We may harness this to create a sequence 
of entangled one-photon wavepackets which are well re- 
solved in time. Let us introduce the following abbrevi- 
ation for a one-photon wavepacket with polarization a 
that has been generated in the j-th generation sequence: 
\a)j = Bl (t J -_ 1 ,T)|0) r (with t = 0). The state after the 
first sequence in more compact form reads: 

|V(i)) = (co|0)i|/o) + c 1 |l) x |/ 1 ))|0) Co |0) cl . (7) 

Suppose we apply a further pulse sequence which recycles 
the atom back to its initial state, i.e., \f a ) — ► |i Q ). Then 
at a time t\ > T we reinitiate the same pulse sequence 
that we have already used previously. It is plausible that 
the wavepackets already generated have in the meantime 
propagated far away from the cavity and thus cannot in- 
fluence the renewed generation sequence. Going through 
the same procedure again we obtain for t > t\ + T: 

m)) = (co|0) 2 |0)x|/o) + ci|l) 2 |l)i|/i))|0) Co |Q) Cl . (8) 

The residual entanglement with the generating system 
can eventually be broken up by making a measurement 
of the internal atomic state in an appropriate basis, e.g., 
|/o) ± For the state in Eq. (||) the resulting reservoir 
state would be one of two states Co|0)i|0)2 ± Ci|l)i|l)2- 
Repeating the generation process n-times followed by a 
final state measurement we produce an n-photon wave 
packet. Note that the description of the reservoir state 
in terms of products of one-photon wavepackets implies 
that the wavepackets can be regarded as independent 
quantum entities. It has to be emphasized that such 
a description is only possible because of the vanishing 
temporal overlap of the individual outgoing wavepackets. 
Actually components of the reservoir state are given by 
products of the operators B^ (tj , T) applied to the multi- 
mode vacuum |0) r . By construction operators belonging 
to different sequences, cf. Eq. (|^), however, commute to 
good approximation: 



[B a (t k ,T),Bl(t j ,T)} = 6 a/3 I dwe^-^\G a {u:,T)\ 2 

~ SapSjk- (9) 

Formally, we may thus regard each creation operator as 
acting on its own vacuum state. Physically, this cor- 
responds to the fact that each wavepacket is contained 
within its private time window of duration T or a box of 
length cT with no overlap between successively generated 
packets, cf. Fig. [|. We have numerically checked the fac- 
torization assumption for a two-photon wavepacket mod- 
eling the reservoir by a discrete set of 1024 "continuum" - 
modes (amounting to more than 10 6 reservoir states) 
embedded in a frequency window of width 40 k c . We 
found that both the Markovian approximation used in 
Eq. ([}]) and the factorization assumption for the two- 
photon spectral density are excellent with relative errors 
of the order of 10~ 3 . 

For any source of entanglement it is essential to 
fathom what the accessible class of states is. In gen- 
eral, a state in a basis spanned by n-qubits is defined 
by N = 2 n+1 — 2 independent coefficients. In our spe- 
cific model the states can be tailored by coupling the 
levels \f a ) by a microwave/Raman pulse inbetween the 
qubit generation sequences. For example, this trans- 
forms Eq. (|) into |V>) = Mo|0)i|/ o ) + Co4|0)i|/i) + 
ad* 1 1) x I fx) - ci d\ 1 1) 1 1 /o)) jO) c 10) C1 • The coefficients d a 
can be chosen at will. With each applied pulse two in- 
dependent parameters are introduced. For n qubits we 
have thus 2n free parameters at hand for the purpose 
of state engineering. Since this is much less than N, 
only a restricted subclass of states can be created. How- 
ever, we emphasize that the accessible class of states in- 
cludes many useful and interesting states. For exam- 
ple MES such as the four Bell states, the GHZ-state 
(| 000) + I 111))/ -\/2 and its higher dimensional counter- 
parts [(\SX, S 2 , S n ) + |1 - Si, 1 - S 2 ,...,l - S n )) / V2, 

Si = 0, 1] can easily be produced. 

The maximum number n of entangled photon 
wavepackets (qubits) our scheme can generate is limited 
by decoherence. Relevant sources of decoherence are: 
(i) laser phase and amplitude fluctuations; (ii) sponta- 
neous emission during the atomic transfer; (iii) absorp- 
tion in the cavity mirrors; (iv) atomic motion. The cho- 
sen configuration minimizes these effects. Stabilization of 
laser phase fluctuations below 1kHz represents a techni- 
cal challenge. In the present scheme, the state produced 
after each cycle only depends on the phase difference be- 
tween the laser beams driving both transitions in Fig. 
1. When these two laser beams are derived from the 
same source the fluctuations in the phase difference are 
effectively suppressed. Amplitude fluctuations cause a 
distortion of the pulse form and lead to incomplete popu- 
lation transfer. An estimate gives that n <C I/AI ~ 10 4 , 
where AI / 1 are the relative intensity fluctuations. Spon- 
taneous emission from the auxiliary levels |r) at rate 
r is quenched by choosing a large detuning \S\ which 
leads to an effective decay rate r c ff = rf2 2 /4<5 2 with 



3 



il = max(f2 Q , g a ). For a peak frequency f^o = 55 Mhz 
and a Gaussian pulse shape for the classical field Vl(t), 
S = 1.5 GHz, g = 55 Mhz, and k c = 50 Mhz one-photon 
pulse durations of around 10 cavity lifetimes are possible. 
Recycling and reinitialization of the medium included a 
conservative estimate would yield a generation rate of 
around 1 MHz. For T = 5MHz the probability of spon- 
taneous emission per cycle is < 10~ 3 . Photon absorption 
in the cavity mirrors is an essential effect for high-Q op- 
tical cavities. In general, it leads to two types of errors: 
Photon absorption and concomitant destruction of the 
entanglement. These errors are evaded by postselection 
through discarding sequences with a number of detected 
photons smaller than n. State distortion jl(| can occur- 
even in the absence of loss of a photon according to: 



l*>= E V*\x) 
xe{o,i}" 



J2 «» 

x£{0,l}™ 



e -( K1 - K0 )Tnx(x)^ ( 10 ) 



where the x are binary representations of different photon 
states, n\{x) is the number of ones contained in x, «o,i 
are the loss rate for modes and 1, respectively. Errors as 
in Eq. ( |Io|) vanish for a cavity with equal absorption rates 
for both polarizations, i.e., kq ~ ki. The optimal way to 
suppress fluctuations induced by the motion of the atom 
is to place the atom at an antinode of both the cavity 
modes and the laser beam (which have to be in standing 
wave configuration) , where the effect of spatial variations 
is minimum and operate in the Lamb-Dicke regime [ [L7J . 
Finally, there might be systematic and random errors in 
the adjustment of the laser pulses used in the recycling 
reinitialization. 




FIG. 2. Ensemble averaged fidelity as a function of the 
number of qubits n. Curves (a),(c), and (e) assume 8 m — 
and e m = 0.0125,0.025,0.1, respectively. Curves (b) and (d) 
are the same as (a) and (c) with S m = 0.05. 

To assess these effects we assume the following imperfect 
mapping in each of the generation sequences: \f a ) — > 
A a \i a ) + B a \f a ), where A a = (1 - e a ) exp(iS a ). In Fig. 
2 we plot the fidelity T(n) of an n-qubit MES produced 
by a source which is ideal exept that the magnitude and 
phase of e a are evenly distributed over a range of [0, e m ] 
and [— e m , e m ]ir, respectively, and the dephasing angle 
is evenly distributed over [— S m ,S m ]Tr. Fig. 2 shows that 
the process is rather robust against global dephasing (S a ) 
but that the correct timing of the 7r-pulses is critical [Q . 
From curve (a) we gather that for errors in the 2% range 



approximately 10 qubits can be created with a fidelity of 
90%. 

We have presented a CQED-based source for the con- 
trolled generation of entangled rt-qubit states where 
the individual qubits are nonoverlapping one-photon 
wavepackets. Our model seems experimentally feasible 
with state-of-the-art equipment and could form the ex- 
perimental basis for multi-party communication in future 
quantum networks. The theory presented can be easily 
adapted to other implementations (e.g. quantum dots 
or single atoms embedded in a host material jl(|) which 
may emerge in the course of time as quantum systems 
with long coherence times. 

Acknowledgements This work was supported by the 
Austrian Research Foundation under grant no. S06514- 
TEC and the European TMR network ERB-FMRX- 
CT96-0087. C.S. thanks Fundacion Andes for support. 



[1] 
[2] 

[3: 

[4 
[5 
[6 
[' 

[«. 
[9 

[io; 



[ii 

[12 
[13 

[14 
[15 
[16 

[17] 



[18] 



D. P. DiVincenzo, Science 270,255 (1995). 

C. H. Bennett, Phys. Today 24 (October 1995) and refer- 
ences cited; A. K. Ekert, Phys. Rev. Lett. 67, 661 (1991). 
J. I. Cirac et al, Phys. Rev. Lett. 78, 3221 (1997); T. 
Pellizzari, ibid. 79, 5242 (1997). 

D. Bouwmeester et al, Nature 390, 575 (1997). 

D. Boschi et al, Phys. Rev. Lett. 80, 1121 (1998). 

Q. A. Turchette et al, Phys. Rev. Lett. 75 4710 (1995). 
G. M. Meyer, H.-J. Briegel, and H. Walther, Europhys. 
Lett. 35, 317 (1997). 

E. Hagley et al, Phys. Rev. Lett. 79, 1 (1997); J. A. Bergou 
and M. Hillary, Phys. Rev. A 55 4585 (1997). 

C. K. Law and H. J. Kimble, J. Mod. Opt. 44 (1997) 2067. 
S. N. Molotkov, and S. S. Nazin, JEPT Lett. 63, 687 
(1996); A. Imamoglu and Y. Yamamoto, Phys. Rev. Lett. 
72, 210 (1994); F. De Martini et al, ibid 76, 900 (1996). 
C. H. Bennett et al, Sci Am. 267(4), 50 (1992); W. Tittel 
et al, |quant-ph/970704S . 

C. H. Bennett et al. , Phys. Rev. Lett. 70, 1895 (1993). 

D. M. Greenberger et al, Am. J. Phys. 58, 1131 (1990); 
D. A. Rice et al, Phys. Lett. A 186, 21 (1994). 

S. N. Molotkov and S. S. Nazin, JEPT Lett. 62, 956 (1995). 
C. W. Gardiner, Quantum Noise (Springer, Berlin, 1991). 
T. Pellizzari et al, Phys. Rev. Lett 75, 3788 (1995); H. 
Mabuchi and P. Zoller, ibid 76, 3108 (1996). 
Here the size of the atomic wavepacket L is smaller than 
the optical wavelength i) = i/i « 1 and the population 
left behind in the wrong level is n 4 /4, which puts a limit 
on the number of photon pulses n <C 4/?7 4 . 
Assuming Gaussian-distributed error |e a | a simple analyt- 
ical estimate yields the exponential decay in Fig. 2. 



4